You are on page 1of 10

Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093

Contents lists available at ScienceDirect

Neuroscience and Biobehavioral Reviews


journal homepage: www.elsevier.com/locate/neubiorev

Review

Interplay of excitation and inhibition elicited by tonal stimulation in pyramidal


neurons of primary auditory cortex
Hisayuki Ojima ∗
Graduate School of Medical and Dental Sciences, Tokyo Medical and Dental University, 1-5-45 Yushima, Bunkyo-ku, Tokyo 113-8749, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Tonal responses of neurons in the primary auditory cortex are a function of frequency, intensity and ear of
Received in revised form 19 October 2010 stimulation. These responses occasionally display suppression. This review discusses how excitatory and
inhibitory synaptic inputs interact to form suppressive responses and how changes in stimulus attributes
Keywords: affect the magnitude and timing of those responses.
Primary auditory cortex Stimulation at the characteristic frequency evokes a stereotyped sequence of depolarization (excita-
Suppression
tory) and then hyperpolarization (inhibitory), as predicted from the canonical circuitry. Some neurons
Excitatory postsynaptic potential
stimulated at higher sound intensities display a prominent increase in the magnitude of hyperpolar-
Inhibitory postsynaptic potential
Non-monotonic rate-level function
ization or a decrease in its latency, both enabling counteraction with the preceding excitation. These
Binaural interaction interactions, in part, underlie the non-monotonic suppression. Furthermore, monaural non-dominant
Spontaneous discharge ear stimulation elicits such a powerful hyperpolarization as to cancel out the depolarization elicited at
Laminar difference dominant ear stimulation, suggesting a linear mechanism for the binaural suppression. Alternatively, it
elicits a depolarization almost equal in magnitude and time course to that elicited at binaural stimulation,
suggesting a nonlinear interaction responsible for the suppression. Laminar differences are also noted
for these inhibitory interactions.
© 2010 Elsevier Ltd. All rights reserved.

Contents

1. Cortical circuitry and inhibitory neurons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2084


2. Involvement of GABAergic interneurons in stimulus specific responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2085
3. EPSP and IPSP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2085
4. Response variability revealed at near-threshold intensity of the CF tones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2086
5. Spontaneous responses as singular or clustered APs and inhibition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2087
6. Forms of suppression observed in AI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2088
7. Suppression in the non-monotonic rate-level function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2088
7.1. Membrane potential changes elicited in non-monotonic responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2088
7.2. Direct measurement of synaptic currents by in vivo whole-cell patch clamp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2089
8. Synaptic activation in the binaural interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2090
9. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2091
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2091

1. Cortical circuitry and inhibitory neurons (Kawaguchi and Kondo, 2002; Markram et al., 2004; Miles, 2000).
Aside from spiny stellate cells, they are biochemically defined to
The neocortical network is comprised of two classes of neu- contain GABA as a neurotransmitter and have been referred to
rons, pyramidal neurons (PNs) and interneurons. Interneurons as inhibitory interneurons, since GABA distances the membrane
are known to consist of heterogeneous cell types with distinct potential of postsynaptic neurons from the threshold for spike
local axon collateralizations and characteristic dendritic arbors generation in adults. More recent studies have revealed the colo-
calization of various neuropeptides with morphologically distinct
GABAergic interneurons. The GABAergic interneurons occur rela-
∗ Tel.: +81 3 5803 5445; fax: +81 3 5803 0186. tively infrequently within the neocortex (approximately 20–25%
E-mail address: yojima.cnb@tmd.ac.jp of the neuronal population; Hendry et al., 1987; Prieto et al., 1994)

0149-7634/$ – see front matter © 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.neubiorev.2010.11.009
H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093 2085

Fig. 1. Schemes of simplified connections of neurons in primary auditory (AI) cortex and the ascending pathway to the AI cortex. (A) As in other primary cortical fields,
neurons in thalamocortical (TC)-recipient zone of the AI cortex receive thalamic afferents from a specific nucleus, the ventral nucleus of the medial geniculate complex (MGC)
as sensory afferents, and association and callosal afferents as extrinsic input (extrinsic). Participants in AI intrinsic connection include recurrent collaterals (recurrent) of
nearby pyramidal neurons (PNs), horizontal collaterals of supragranular PNs (pt) located at remote loci within the AI (intrinsic; Ojima et al., 1991, 1992), and collaterals of
local interneurons. Note that a minicolumn has a horizontal width of 50–75 ␮m (Winer, 1984) and the horizontal extent of a TC afferent is 250 ␮m in diameter (Hashikawa
et al., 1995; Cetas et al., 1999). Pyramidal neurons, triangles; GABAergic interneurons, ovals; TC neurons, circles; and extrinsic neuron, square. (B) There are three major
commissural projections in the auditory ascending pathway; at the level of the cochlear nucleus, the inferior colliculus, and cortex (corpus callosum).

and the range of their proportions in different layers is relatively Synaptic transmission to the cortical recipient neurons is highly
small (15–30% Hendry et al., 1987), except for layer 1 in which reliable (Rose and Metherate, 2005), and this may be due to the
more than 90% of the neurons are GABAergic. Despite their rela- preferential positioning of TC synapses at locations near the soma
tively infrequent occurrence, the GABAergic interneurons play a (Richardson et al., 2009). Response properties of AI neurons are
very important role in the regulation of neuronal excitability, con- generated by simple inheritance of thalamic activities, linear con-
struction of receptive field properties, temporal precision of spike struction from simple to complex receptive fields, or nonlinear
timing, and signal synchronization leading to the generation of cor- assembly of distinct activities (Creutzfeldt et al., 1980; Miller et al.,
tical rhythms (Bacci and Huguenard, 2006; Creutzfeldt et al., 1974; 2001). However, it is also very likely that they depend on cortical
Ferstet and Koch, 1987; Fricker and Miles, 2001; Oswald et al., 2009; circuitry involving diverse types of interneurons (McMullen and
Sadagopan and Wang, 2010; but see Priebe and Ferster, 2008). Glaser, 1982; Prieto et al., 1994), just as in the model proposed for
Excitability of neurons in a small cortical domain or column is the primary visual field in which interactions of inhibition with
determined by afferents of multiple sources such as thalamocor- excitation serve to refine various stimulus selectivities (Ferstet and
tical (TC) neurons, cortical interneurons, neighboring PNs, distant Koch, 1987; Shapley et al., 2003). In this review, by relating spike
but within-field PNs, and association and callosal neurons (Fig. 1). rate changes to behaviors of membrane potentials, I propose pos-
Response properties of sensory cortical neurons in the primary sible wiring schemes which may account for the responses derived
recipient zone of TC afferents are largely determined by a network from interactions of excitation and inhibition evoked by tonal stim-
involving feedforward activation of pyramidal/spiny stellate neu- ulation, usually at the neuron’s characteristic frequency (CF).
rons and GABAergic interneurons, known as canonical circuitry.
This model was originally proposed for the visual cortex, includ- 3. EPSP and IPSP
ing both supragranular and infragranular microcircuits (Douglas
and Martin, 1991, 2004), but is likely to be also applicable at least A stimulus-evoked hyperpolarized deflection of membrane
to supragranular layer in the auditory cortex, (Cruikshank et al., potential (inhibitory postsynaptic potential, IPSP) following a depo-
2002; Shu et al., 2003; Rose and Metherate, 2005; Richardson et al., larized membrane potential (excitatory postsynaptic potential,
2009). EPSP) at the onset of acoustic stimulation was revealed by elec-
trophysiological recordings of membrane potentials from single AI
2. Involvement of GABAergic interneurons in stimulus neurons with conventional sharp glass-microelectrodes (Ojima and
specific responses Murakami, 2002; De Ribaupierre et al., 1972; Volkov and Galazjuk,
1991; Volkov and Galazyuk, 1992). A similar sequence of EPSP
Unlike the primary visual field, typical spiny stellate neurons are and IPSP was also induced by electrical stimulation of the MGC
very rare in layer 4 of the primary auditory (AI) field (Fitzpatrick and (Mitani et al., 1985). GABA agonist/antagonist application directly
Henson, 1994; McMullen and de Venecia, 1993; Meyer et al., 1989; into AI dramatically changed the frequency response field and
Smith and Populin, 2001; Winer, 1984), and PNs in layer 4 and lower frequency-intensity function both in vivo (Chen and Jen, 2000; Kaur
layer 3 are regarded as the major recipients of TC afferents in the et al., 2004; Wang et al., 2000, 2002) and in vitro (Broicher et al.,
auditory system (Hashikawa et al., 1995; Huang and Winer, 2000; 2010). Beyond the stimulus-driven induction of inhibitory activ-
McMullen and de Venecia, 1993; Smith and Populin, 2001). ity or its anti-epileptic effect, findings have accumulated as to how
In the auditory canonical circuitry, TC afferents are derived inhibitory inputs serve to shape acoustic receptive fields (Liu et al.,
from the ventral nucleus of the medial geniculate complex (vMGC). 2007; Ojima and Murakami, 2002; Tan et al., 2004, 2007; Tan and
2086 H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093

2010a,b; Ojima and Murakami, 2002). Thus, it is predicted that


the interactions of excitatory and inhibitory synaptic inputs to PNs
result in diverse patterns in different layers.

4. Response variability revealed at near-threshold intensity


of the CF tones

Excitability of AI neurons can be demonstrated in the tuning


curve constructed on a frequency and intensity matrix from which
an optimal stimulus frequency is defined as the CF at the mini-
mum threshold intensity (Schreiner, 1992). It was observed that,
at or slightly above the threshold, spike generation was not fully
secured but fluctuated on a trial-by-trial basis. Namely, in repeated
presentation of the same CF tone, some trials succeeded in evoking
spikes in AI PNs, but other trials failed. If this fluctuation in fir-
ing is examined in terms of membrane potential changes, we will
see occurrence of an EPSP either with or without action potentials
(EPSP + APs or subthreshold EPSP) at the onset of acoustic stim-
ulation and, possibly, after a short delay, an IPSP. The canonical
circuitry suggests that the critical determinant of IPSP generation in
PNs is whether or not the activation of cortical GABAergic interneu-
rons by TC afferents is suprathreshold.
The cortical network may be more complicated than that repre-
sented by the canonical circuitry. Depending on whether responses
of the cortical PN and interneuron in the canonical circuitry
cross the threshold or remain below, four different patterns of
excitation–inhibition sequence can be expected for individual PNs.
Fig. 2. Canonical circuitry and possible response sequences of excitation and inhi-
bition predicted from the circuitry. (A) Scheme explaining the canonical circuitry in Thus, suprathreshold EPSP (supraEPSP) elicited both in the PN and
which TC neurons in the ventral nucleus of the MGC sends feedfoward projections interneuron will yield a typical membrane potential sequence start-
to PNs (open triangle) as well as interneurons (filled circle), which in turn send ing with an EPSP + AP followed by an IPSP (type I sequence) in this
inhibitory projections to the PNs. (B) Different sequences of excitation and inhi-
PN. Similarly, supraEPSP in the PN and subthreshold EPSP (subEPSP)
bition (types I–IV) are predicted in the pyramidal neurons depending on whether
activation of the pyramidal neurons and interneurons of the canonical circuitry is
in the interneuron will yield an EPSP + AP only (type II sequence) in
suprathreshold or subthreshold (see A for possible combinations of suprathresh- the PN. Moreover, subEPSP in the PN and supraEPSP in the interneu-
old and subthreshold responses). (C) Stimulation with the neuron’s characteristic ron will yield a subEPSP followed by an IPSP (type III sequence), and
frequency at a fixed intensity slightly above threshold for spike generation elic- subEPSP elicited both in the PN and interneuron will yield a subEPSP
its fluctuated responses that are suprathreshold in some trials and subthreshold
alone (type IV sequence) (see Fig. 2B).
in other trials. This fluctuation is represented by differently colored postsynaptic
membrane potentials elicited. Each neuron elicits only two types of the response Each of these 4 response types would be potentially available
sequences shown schematically in B. Combination A, type I or type IV. Combination for every neuron. However, for responses to many repeats of a
B, type I or type III. Combination C, type II or type IV. Combination of type II and type stimulus tone with the same CF at the intensity slightly above
III is never observed. Each trace is a single trial. Vertical think bars indicate 5 mV.
threshold, only two types of sequences occurred in each neuron
Horizontal thick bars indicate the stimulus period of 50 ms.
(Fig. 2C, unpublished observation). Many AI neurons (16/40) dis-
played either a type IV or I sequence in the fluctuated response
(combination A). The majority of neurons (20/40) showed either
Wehr, 2009; Wehr and Zador, 2003; Wu et al., 2006, 2008), how a type III or I sequence (combination B). Nearly 10% of PNs (4/40)
they define direction selectivity to frequency-sweep sounds (Ye exhibited either a type II or IV sequence (combination C). We never
et al., 2010; Zhang et al., 2003), how they regulate temporal dis- observed PNs that showed other possible combinations such as
charge patterns (Huetz et al., 2009; Wehr and Zador, 2005), and type II and type III sequence as the fluctuated responses. Consider-
how they are involved in the response properties related to or ing that stronger neuronal activation is needed for the generation
assumed to be related to the spatial localization of a sound source of an supraEPSP (in type II sequence) than of an subEPSP (in type
(Chadderton et al., 2009; Ojima and Murakami, 2002). III sequence), this combination indeed may not be possible unless
EPSP and IPSP recorded from PNs at the stimulus onset, espe- we assume GABAergic interneurons with the non-monotonic rate-
cially in the TC recipient zone (layer 4 and lower layer 3), are elicited level function to explain the absence of an IPSP following the
chiefly by activation of excitatory input from the thalamus and supraEPSP of type II sequence. The non-random arrangement of
inhibitory input from the cortex. These oppositely deflected post- these patterns of synaptic sequences is presumably the reflection
synaptic potentials (PSPs) show variability in their magnitude and of some anatomical constraints in association with synaptic inputs
timing according to the varying stimulus attributes. Canonical cir- to individual PNs, interneurons or both.
cuitry (Fig. 2A) has been proposed as the basic network underlying Each layer has a distinct local network. Indeed, the diverse pat-
these interactions (Douglas and Martin, 1991, 2004; Liu et al., 2007; terns of synaptic sequences in individual PNs can be ascribed in part
Oswald et al., 2006; Zhou et al., 2010). Its simplified connectivity, to their different laminar localizations. By iontophoretically inject-
however, does not explain the full range of response diversities of ing a dye filled in the glass-microelectrode used for intracellular
AI neurons (Liu et al., 2007). Especially, since the distribution of recording, neurons can be labeled for their laminar locations. This
TC afferents is heterogeneous across cortical supragranular layers showed that almost all PNs recorded in layer 2 showed a type II or IV
(Hashikawa et al., 1995; Huang and Winer, 2000; McMullen and sequence (Ojima and Murakami, 2002). Their onset EPSP (onEPSP),
de Venecia, 1993), response diversities are likely to vary signifi- irrespective of AP generation or failure, is not followed by IPSP. This
cantly between neurons in different layers (Atencio and Schreiner, implies that the canonical-like circuitry is not the dominant type
H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093 2087

Fig. 3. Relationship between the number of spikes and the IPSP amplitude following them. Membrane potential changes in response to repeated presentation (stimulus
interval, 1.1 s) of a single stimulus tone with the neuron’s characteristic frequency at 10 dB above the minimal threshold intensity (left) and those occurring spontaneously
(right). Uppermost traces show superimposition of all trials, with prominently larger hyperpolarizations pointed by small arrows. Note mixture of traces with small or large
amplitudes of IPSPs in both stimulated and spontaneous cases. The IPSPs with larger amplitudes are almost always preceded by a burst of action potentials, i.e., triple spikes
(lower left, red traces) for evoked firing, or double or more-than-double spikes for spontaneous firing (lower right, red traces). Horizontal lines in superimposed and individual
traces indicate resting membrane potential levels.

of connection in layer 2 and, probably, upper layer 3. Such differ- they were clustered (Fig. 3, unpublished observation). Single PNs
ences in local synaptic interactions may also be related to laminar in cat AI have a rich plexus of local collaterals with an extent of
diversity of dynamic sound processing of PNs in cat AI (Atencio and approximately 500 ␮m in diameter around the cell of origin (Ojima
Schreiner, 2010a,b; Huggenberger et al., 2009; Sakata and Harris, et al., 1991, 1992). These collaterals are most likely to make synaptic
2009). contacts with nearby neurons including GABAergic interneurons,
which in turn make inhibitory synaptic contacts with the PNs from
5. Spontaneous responses as singular or clustered APs and which the axon collaterals are derived. Data from the auditory cor-
inhibition tex are limited (Atzori et al., 2001; Barbour and Callaway, 2008;
Oswald and Reyes, 2008; Oswald et al., 2009), but, in many sen-
In addition to stimulus-driven APs, chiefly via activation of sory systems, recurrent network is formed both between different
TC input, APs also occur spontaneously in a singular or bursting PNs and between PNs and interneurons (Holmgren et al., 2003;
mode (Eggermont, 1992; Valentine and Eggermont, 2001). It was Sanchez-Vives and McCormick, 2000; Staiger et al., 2002, 2009;
observed in some neurons that no IPSP followed spontaneous APs Thomson and Lamy, 2007). This raises the possibility that APs
when they were singular (De Ribaupierre et al., 1972; Tan et al., evoked spontaneously in a presynaptic PN, when they are singular,
2004), but that an IPSP frequently followed spontaneous APs when are not sufficient to bring the membrane potential of postsynap-
2088 H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093

Fig. 4. Two representative response patterns of suppression found in the non-monotonic amplitude-versus-level function. Stimulus periods are shown by thick lines on time
axis. (A) As sound intensity increases from 35 to 70 dB SPL, amplitude of the onEPSP decreases in association with its decreasing duration, probably caused by increasingly
earlier onset of IPSP as tone intensity increases (short arrows). Individual traces represent the average of 10 repeated responses. Characteristic frequency of this pyramidal
neuron is 14.3 kHz. (B) As sound intensity increases from 10 to 80 dB SPL, amplitude of the onEPSP, together with decreasing firing rate (shown in parentheses), decreases in
association with the increasing amplitude of the IPSP following it. Characteristic frequency of this pyramidal neuron is 20.1 kHz. From Ojima and Murakami (2002).

tic interneurons above threshold. However, when spontaneous APs and stimulus-driven discharges is also known to occur during self-
occur in the PN more than once during a narrow time window (i.e., a vocalization of monkeys (Eliades and Wang, 2003; Müller-Preuss
burst), probably owing to the temporal summation of EPSP elicited and Ploog, 1981).
by each of the APs of a burst, these APs easily bring membrane How EPSP and IPSP interact to suppress the responses
potentials of the postsynaptic interneurons above threshold, thus of AI neurons was systematically examined by intracellularly
entailing AP generation in these interneurons. The APs generated in recording membrane potentials using conventional sharp glass-
the GABAergic interneurons in turn elicit IPSPs in the presynaptic microelectrodes (Ojima and Murakami, 2002). The results revealed
PN in a reciprocal manner. counteractions of EPSP and IPSP in a stimulus-dependent manner
Reduced spike count immediately after burst-like click trains by virtue of the disproportionate growth of IPSP, advanced onset of
has been reported in vivo (Eggermont, 1992), although it is not IPSP, and suppressive influences of non-dominant ear stimulation
known how the changes in firing mode of AI neurons are related on dominant ear stimulation.
to sound processing. Intracortical electrical stimulation at high
rates increased the occurrence of bursting activities and led to
7. Suppression in the non-monotonic rate-level function
changes in cross-correlation between neuronal pairs (Valentine
and Eggermont, 2003). It also accompanied a shift in the CFs of
7.1. Membrane potential changes elicited in non-monotonic
neurons at the stimulation site. Thus, elevated activity at a certain
responses
cortical site may alter the synchronization pattern of cortical neu-
rons, presumably necessary for the changed perception of acoustic
At least two types of interactions of EPSP and IPSP have been
attributes. It is conceivable that the occurrence of strong inhibition
elucidated for neurons showing the non-monotonic rate-level
following burst firing may be involved in changes in the neuronal
function of spike discharge. The first interaction underlying the
synchronization.
reduced spike discharge (Fig. 4A) is that the onset latency of the
IPSP becomes shortened to a much greater extent than that of the
6. Forms of suppression observed in AI preceding onEPSP as sound intensity increases. Because of this ear-
lier onset of IPSP, the peak/falling phase of the preceding EPSP is
Suppression, reflected by reduced spike count, is known to chipped off, leading eventually to a shorter duration and decreased
shape various forms of response properties of auditory cortical amplitude. Thus, at higher sound intensities, EPSP and IPSP come
neurons. Representative suppressions in the AI cortex include: (1) to counteract in a summative manner (Eccles, 1968), since they
lateral side-band suppression, i.e., reduction in firing rate at the become temporally coincident. The shortened duration of onEPSP
neuron’s CF tone by prior or simultaneous presentation of an addi- also leads to a reduced jitter in the onset timing of evoked APs or a
tional tone at different frequencies (Brosch and Schreiner, 1997; more consistent delay of the AP onset across trials. This is due to a
Calford and Semple, 1995; Sutter and Schreiner, 1995; Sutter et al., narrower time window for the onEPSP to become suprathreshold.
1999); (2) suppression in non-monotonic rate versus level func- In another type (Fig. 4B), as sound intensity increases, the IPSP peak
tion, i.e., reduction in firing rate at higher sound intensities (Brugge amplitude increases greatly to such a degree that the preceding
et al., 1969; Heil et al., 1994; Pfingst and O’Connor, 1981; Phillips onEPSP amplitude becomes substantially reduced. Here, delay time
and Cynader, 1985; Phillips and Irvine, 1981; Suga, 1977; Sutter of the IPSP appears to contribute little to the reduced amplitude of
and Schreiner, 1995); and (3) suppression in binaural interaction, the preceding onEPSP. This results in a decreased spike count.
i.e., reduction in firing rate at stimulation of both ears relative to These types of interaction of EPSP and IPSP underlying the non-
stimulation of one ear alone (Benson and Teas, 1976; Brugge et al., monotonic firing pattern are typically observed for PNs located in
1969; Brugge and Merzenich, 1973; Calford and Semple, 1995; the TC recipient zone of the AI cortex. Neurons in layer 2, however,
Hall and Goldstein, 1968; Imig and Adrián, 1977; Imig and Brugge, show different membrane potential behaviors. EPSP is elicited in
1978; Imig and Reale, 1981; Kitzes et al., 1980; Middlebrooks et al., layer 2 PNs but is not eventually followed by IPSP across a wide
1980; Phillips and Cynader, 1985; Phillips et al., 1985; Phillips and range of sound intensities at their CF and frequencies other than CF.
Irvine, 1983; Semple and Kitzes, 1985). Suppression of spontaneous Nevertheless, onEPSP recorded from layer 2 PNs changes its ampli-
H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093 2089

Fig. 5. Two representative response patterns of suppression found in the binaural interaction. (A) As shown in the upper panel, binaural stimulation (BIN) elicits a smaller
amplitude of an onEPSP, together with fewest spikes, when compared to an onEPSP amplitude elicited by the dominant (dom, IPSI) ear stimulation. The binaural suppression
of the onEPSP can be explained chiefly by linear summation of the response elicited at independent stimulation of each ear (i.e., IPSI dominant and CONTRA non-dominant
ear). For this neuron, as shown in the lower panel, non-dominant contralateral stimulation consistently elicits a strong IPSP at a wide range of stimulus intensities (20–80 dB
SPL). (B) Binaural stimulation elicits an onEPSP whose amplitude is almost equal to that elicited at non-dominant ear stimulation (non-dom, CONTRA) but much smaller
than that elicited at dominant ear stimulation (dom, IPSI). Simple summation of the responses evoked by independently stimulating each ear cannot account for the changed
amplitude of both EPSP and the following IPSP. Individual traces represent the average of 10 repeated trials. Arrowheads indicate resting membrane potential levels. Thick
horizontal lines show stimulus duration. Asterisks in A point to two hyperpolarizing deflection of the membrane potential. From Ojima and Murakami (2002).

tude in a non-monotonic manner. Thus, the onEPSP amplitude depending on the firing rate, although they do not display the best
decreases without accompanying an IPSP even as sound inten- intensity. Furthermore, without a systematic amplitude map across
sity increases (see Fig. 6, Ojima and Murakami, 2002). The absence the auditory cortex, how do non-monotonic neurons distinguish
of IPSP following the onEPSP can possibly be explained by differ- two intensity levels at which an equal magnitude of discharges
ences in the local networks in layer 2 and the recipient zone of TC is evoked, one lower and the other higher than the neuron’s best
afferents. Physiologically, PNs in the recipient zone are monosy- intensity? So far, there has been little evidence for a systematic
naptically driven by stimulation of TC afferent (Cruikshank et al., cortical amplitude map except in bats (Suga, 1977). Monotonic and
2002; Metherate and Cruikshank, 1999; Mitani and Shimokouchi, non-monotonic domains seem to be rather topographically orga-
1985; Rose and Metherate, 2001; Volkov and Galazjuk, 1991). It nized without forming an intensity map in the cat AI (Schreiner,
is known in AI that stimulation of the TC-recipient zone evokes 1992; Sutter and Schreiner, 1995).
EPSP in layer 2 PNs (Barbour and Callaway, 2008) as in motor
cortex (Kang et al., 1994). Thus, layer 2 PNs are likely to be a
7.2. Direct measurement of synaptic currents by in vivo
disynaptic target of TC afferents (Mitani et al., 1985), conforming
whole-cell patch clamp
to the findings in the visual system (Toyama et al., 1974; Ferster
and Lindström, 1983). Since a rich axon plexus from TC-recipient
Membrane potential measurements by conventional sharp
layer 3 PNs extends into layer 2 (Ojima et al., 1991, 1992), the
micro-electrodes in vivo have elucidated voltage changes caused
above findings suggest that tonally driven excitation carried by
by a net current as the consequence of linear summation of inward
TC afferents can reach layer 2 neurons polysynaptically via exci-
and outward currents (EPSC and IPSC) passing through respec-
tatory synapses of layer 3 PNs. A plausible interpretation is that
tive ligand-gated channels. However, the currents are not directly
these polysynaptic excitatory inputs cannot drive local interneu-
measured with this technique. Recent in vivo patch-clamp tech-
rons that are presynaptic to layer 2 PNs or bring their membrane
nique has allowed such measurements in the whole-cell voltage
potential only to a subthreshold level. The canonical circuitry may
clamp mode (Wehr and Zador, 2003; Zhang et al., 2003). It can
not be applicable here. Therefore, the non-monotonicity found in
separately measure excitatory and inhibitory currents by clamp-
layer 2 PNs would be inherited from cortical layer 3 PNs that are also
ing the membrane potential close to the equilibrium potentials for
non-monotonic in the their rate-level function. It can be implied
inhibitory and excitatory inputs, respectively, thus making it pos-
that connections would be formed between PNs located in differ-
sible to measure TC excitatory and cortical inhibitory contribution
ent layers of a column if their rate-level function is of the same
independently.
type.
The use of this technique has revealed that acoustically evoked
Are neurons having the best intensity defined from the non-
IPSC and EPSC are rather stereotyped in terms of their timing and
monotonic rate-level function the neuronal correlates for intensity
relative magnitude. Namely, an IPSC follows an onset EPSC with a
coding or loudness perception? On the basis of the findings
relatively constant delay time across a wide range of stimulus fre-
that increases in the neuronal proportion of non-monotonic type
quencies and intensities, implicating that the EPSC and IPSC interact
occur only in animals trained for loudness discrimination (Polley
minimally. Besides, although excitatory and inhibitory conduc-
et al., 2004, 2006), it is proposed that non-monotonic neurons are
tances vary widely according to the stimulus parameters, with the
involved in intensity encoding in such a way that they may change
maximum at the neuron’s CF, the ratio between them remains rel-
firing rates as a population in proportion to the increase in sound
atively constant. It indicates that they are balanced; namely, the
intensity. However, it seems premature to say that non-monotonic
contribution of inhibition to the creation of the response non-
neuronal population is solely responsible for loudness perception.
monotonicity is limited. In AI, therefore, the frequency tuning and
Monotonic neurons are presumably able to encode sound levels
intensity tuning might be generated without the interactions of
2090 H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093

excitation and inhibition (Tan and Wehr, 2009; Wehr and Zador, ulation. However, the amplitude and latency of these PSPs varied
2003), implying that tuning properties of excitatory and inhibitory considerably depending on which ear was stimulated (Ojima and
inputs of AI neurons are inherited from the presynaptic TC neurons Murakami, 2002). Some responses at binaural stimulation, nev-
in the vMGC. ertheless, can be explained by the linear summation of the PSPs
However, in later studies, two forms of interaction between elicited by independent stimulation of each ear. As exemplified
EPSCs and IPSCs have been noted. EPSC and IPSC are unbalanced in by an EI neuron in Fig. 5A, monaural, dominant ear stimulation
a subset of AI neurons. These interactions provide direct evidence evoked a typical sequence of the PSPs (Fig. 5A, upper). Monaural,
for the involvement of inhibition in cancelling out of the preceding non-dominant ear stimulation evoked an IPSP alone without a pre-
excitation according to a non-monotonic rate-level function, as pre- ceding EPSP across a wide frequency range (Fig. 5A, lower). This
viously suggested by the sharp microelectrode measurements. As IPSP coincided with the EPSP evoked by the dominant, opposite ear
tone intensity increases, IPSC increases disproportionately and/or stimulation. Therefore, if spatial and temporal summation holds
its onset becomes much earlier compared to that elicited at true, simultaneous stimulation of two ears (binaural stimulation)
weaker sound intensities. Higher sound intensities thus lead to the predicts a decreased level of the evoked EPSP, leading to decreased
shortening of a delay time between the EPSC and IPSC, allowing discharges representing the EI type of binaural interaction.
counteraction of the onset excitation with the inhibition following In some EI neurons, however, linear summation cannot ade-
it (Tan et al., 2007; Wu et al., 2008). quately explain the reduction in onEPSP amplitude. Fig. 5B shows
that the onEPSP amplitude at binaural stimulation is almost equal
to that at monaural, non-dominant ear stimulation. The EPSP at
8. Synaptic activation in the binaural interaction dominant ear stimulation, by definition, is larger in amplitude than
that at the non-dominant ear stimulation, and therefore larger than
In the visual system, callosal axons projecting through the that at the binaural stimulation. Thus, if linear summation under-
corpus callosum (callosal projections) to the contralateral hemi- pins the binaural interaction, it would be predicted that the sum
sphere terminate in a restricted zone bordering between areas of the non-dominant and dominant ear responses be larger than
17 and 18 (Fisken et al., 1975; Harvey, 1980; Shatz, 1977). In either of the component monaural responses. In reality, however,
contrast, terminal axons of auditory callosal projections are exten- this summation representing the binaural response does not yield
sively distributed across AI (Code and Winer, 1986; Imig and such increased responses.
Brugge, 1978; Rüttgers et al., 1990; Wallace and Harper, 1997). Alternative interpretation might adopt the mechanisms of
However, they are not homogenously distributed but constitute suppression underlying the non-monotonic rate-level function.
multiple bands across the AI cortex in cats and ferrets (Imig and Because the auditory ascending pathway is basically bilateral, bin-
Brugge, 1978; Imig and Reale, 1981; but see Brandner and Redies, aural stimulation is likely to recruit more projections, thereby
1990). It is thought that these callosal bands are interleaved with driving cortical neurons more intensively, than monaural stimu-
zones containing predominantly terminals of association neurons. lation. Binaural stimulation therefore is comparable to stimulation
The callosal and association zones are assumed to be function- at higher sound intensities (Semple and Kitzes, 1993a,b; Zhang et
ally different and referred to as, respectively, EE and EI binaural al., 2004). Namely, the reduced amplitude of onEPSP at binaural
interaction bands. In EE bands, neurons show facilitation at binau- stimulation would be a consequence of more intensive stimulation
ral stimulation (i.e., binaural response magnitude is greater than of non-monotonic neurons than at monaural stimulation. However,
monaural response magnitude), while in EI bands, neurons show this does not seem to be the case. According to one of the mech-
suppression at binaural stimulation (i.e., response magnitude at anisms underlying the non-monotonic rate-level function, an IPSP
dominant ear stimulation is greater than that at binaural stim- becomes elicited so earlier as to counteract the preceding EPSP at
ulation). Some neurons display responses only to stimulation of higher sound intensities (Ojima and Murakami, 2002; Wu et al.,
one ear (EO binaural interaction) and seem to be dispersed across 2006). However, as shown in Fig. 5B, there is no such shorten-
the AI cortex without constituting distinct bands (Imig and Brugge, ing of the onset latency of the IPSP at the binaural stimulation
1978). compared to that at the non-dominant ear stimulation. Similarly,
It should be noted that the binaural interaction does not sim- according to another mechanism, an IPSP becomes augmented to
ply represent the interaction between ipsilateral and contralateral a level sufficient to unbalance the preceding EPSP. However, no
acoustic inputs per se. In the auditory ascending pathway, because such growth of the IPSP is associated with the onEPSP at the binau-
of the commissural projections at two subcortical levels (the ral stimulation. Thus, the underlying mechanisms are different and
cochlear nucleus and the inferior colliculus), acoustic stimulation likely include some nonlinear interactions. These remain to be yet
to one ear eventually induces bilateral activation at any level. examined.
Since each hemisphere receives callosal projections from the oppo- Recordings from most of the so-called EO PNs showed no
site hemisphere (anatomically; Code and Winer, 1986; Imig and substantial differences in the magnitude and latency of onEPSP
Brugge, 1978; Rüttgers et al., 1990; Wallace and Harper, 1997) between the dominant ear stimulation and the binaural stimula-
(physiologically; Goycoolea et al., 2005; Kitzes and Doherty, 1994; tion. Shaping of responses by inhibition is likely to be minimal,
Mitani and Shimokouchi, 1985), it is highly possible that monaural although some EO neurons had an IPSP following an onEPSP. Since
stimulation alone can drive cortical neurons in each hemisphere. the responses evoked by binaural stimulation do not differ from
This is by way of TC projections ipsilaterally and callosal inputs those evoked by dominant ear stimulation, at least two patterns of
contralaterally. Therefore, elucidating the binaural interactions is responses to the non-dominant ear stimulation could be assumed.
tantamount to characterizing the composite PSPs elicited by con- In one pattern, non-dominant ear stimulation would generate null
currently activated TC and callosal inputs to PNs and to examine onset response; neither EPSP nor IPSP is evoked at the stimulus
how different the composite responses are among different aural onset. In the other pattern, non-dominant ear responses would
stimulation styles. Also, ‘binaural stimulation’ generally means the be temporally lagged behind dominant ear responses to such an
bilateral presentation of a tone at the same frequency, intensity, and extent that these two responses could not coincide; the possibility
phase. Thus, it may correspond to stimulation at the zero azimuthal of the temporal interaction between the two responses is minimal
point in the free field. in this pattern. In reality, with the exception of a single neuron,
Intracellular recording of PSPs typically showed a stereotyped no membrane potential change was evoked by non-dominant ear
sequence of an onEPSP followed by an IPSP at any style of aural stim- stimulation (14/15, unpublished observation). Thus, like some EI
H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093 2091

neurons, EO binaural interaction can simply be explained by the Calford, M.B., Semple, M.N., 1995. Monaural inhibition in cat auditory cortex. J.
linear summation of PSP evoked separately by stimulation of each Neurophysiol. 73, 1876–1891.
Cetas, J.S., de Venecia, R.K., McMullen, N.T., 1999. Thalamocortical afferents of
ear alone. Lorente de Nó: medial geniculate axons that project to primary auditory cortex
Although neurons showing the same binaural interaction prop- have collateral branches to layer I. Brain Res. 830, 203–208.
erties tend to be arranged in a columnar fashion in the AI cortex Chadderton, P., Agapiou, J.P., McAlpine, D., Wargrie, T.W., 2009. The synaptic
representation of sound source location in auditory cortex. J. Neurosci. 29,
(Imig and Adrián, 1977), the roles that neurons with different bin- 14127–14135.
aural interaction properties play in the localization of a sound Chen, Q.C., Jen, P.H., 2000. Bicuculline application affects discharge patterns, rate-
source are little understood. An important question is how inhibi- intensity functions, and frequency tuning characteristics of bat auditory cortical
neurons. Hear Res. 150, 161–174.
tion grows or how it interacts with the preceding EPSPs if sound
Creutzfeldt, O., Hellweg, F.C., Schreiner, C., 1980. Thalamocortical transformation of
source location shifts in a systematic manner, or more specifi- responses to complex auditory stimuli. Exp Brain Res. 39, 87–104.
cally, in the systematic changing of the interaural time or level Creutzfeldt, O.D., Kuhnt, U., Benevento, L.A., 1974. An intracellular analysis of visual
cortical neurones to moving stimuli: response in a co-operative neuronal net-
difference. This may not be available at present, but recent in vivo
work. Exp. Brain Res. 21, 251–274.
measurements of PSPs evoked by a noise emitted from various spa- Cruikshank, S.J., Rose, H.J., Metherate, R., 2002. Auditory thalamocortical synaptic
tial locations revealed the occurrence of subEPSP with substantial transmission in vitro. J. Neurophysiol. 87, 361–384.
amplitudes even at locations much deviated from the neuron’s pre- Code, R.A., Winer, J.A., 1986. Columnar organization and reciprocity of commissural
connections in cat primary auditory cortex (AI). Hear Res. 23, 205–222.
ferred location where the EPSP amplitude (therefore, the firing rate Douglas, R.J., Martin, K.A., 2004. Neuronal circuits of the neocortex. Annu. Rev. Neu-
of APs) was maximal (Chadderton et al., 2009). It also suggested the rosci. 27, 419–451.
possible involvement of the steepness of the onEPSP rising phase Douglas, R.J., Martin, K.A., 1991. A functional microcircuit for cat visual cortex. J.
Physiol. 440, 735–769.
in the detection of the sound source by cortical neurons. Further De Ribaupierre, F., Goldstein Jr., M.H., Yeni-Komshian, G., 1972. Intracellular study
studies at the synaptic level will be awaited. of the cat’s primary auditory cortex. Brain Res. 48, 185–204.
Eccles, J.C., 1968. The Physiology of Nerve Cells. Johns Hopkins University Press,
Baltimore.
9. Conclusions Eliades, S.J., Wang, X., 2003. Sensory–motor interaction in the primate
auditory cortex during self-initiated vocalizations. J. Neurophysiol. 89,
2194–2207.
Synaptic mechanisms underlying the spike rate changes evoked Eggermont, J.J., 1992. Stimulus induced and spontaneous rhythmic firing of single
by changing acoustic attributes were originally explored by units in cat primary auditory cortex. Hear Res. 61, 1–11.
Ferstet, D., Koch, C., 1987. Neuronal connections underlying ori-
measuring membrane potentials in vivo. The depolarizing and entation selectivity in cat visual cortex. TrendsNeurosci. 10,
hyperpolarizing shifts in membrane potential have revealed the 487–492.
importance of inhibitory synaptic input in shaping the spatiotem- Ferster, D., Lindström, S., 1983. An intracellular analysis of geniculo-cortical connec-
tivity in area 17 of the cat. J. Physiol. 342, 181–215.
poral response field of single PNs. Direct in vivo measurements Fitzpatrick, D.C., Henson Jr., O.W., 1994. Cell types in the mustached bat auditory
of synaptic currents have refined these mechanisms and clarified cortex. Brain Behav. Evol. 43, 79–91.
the mechanisms for precise interactions of the strength and tim- Fisken, R.A., Garey, L.J., Powell, T.P., 1975. The intrinsic, association and commissural
connections of area 17 on the visual cortex. Philos. Trans. R. Soc. Lond. B: Biol.
ing of excitatory/inhibitory synaptic conductance in the receptive
Sci. 272, 487–536.
field shaping. This approach has been extended even to suggest Fricker, D., Miles, R., 2001. Interneurons, spike timing, and perception. Neuron 32,
non-linear interactions of excitation and inhibition (Machens et al., 771–774.
Goycoolea, M., Mena, I., Neubauer, S., 2005. Functional studies of the human audi-
2004), importance of the spike generation mechanism (Ye et al.,
tory pathway after monaural stimulation with pure tones. Establishing a normal
2010), and laminar specificity of processing of acoustic signals database. Acta Otolaryngol. 125, 513–519.
(Zhou et al., 2010) in the auditory system. Auditory cortex seems to Hall 2nd, J.L., Goldstein Jr., M.H., 1968. Representation of binaural stimuli by single
be influenced by non-auditory signals or even by the brain internal units in primary auditory cortex of unanesthetized cats. J. Acoust. Soc. Am. 43,
456–461.
state (Ojima et al., 2010). One of the ultimate goals in the auditory Harvey, A.R., 1980. A physiological analysis of subcortical and commissural projec-
sensory physiology is to describe the interactions of sensory infor- tions of areas 17 and 18 of the cat. J. Physiol. 302, 507–534.
mation and memory-based internal information at the synaptic Hashikawa, T., Molinari, M., Rausell, E., Jones, E.G., 1995. Patchy and laminar termi-
nations of medial geniculate axons in monkey auditory cortex. J. Comp. Neurol.
level. 362, 195–208.
Heil, P., Rajan, R., Irvine, D.R., 1994. Topographic representation of tone inten-
sity along the isofrequency axis of cat primary auditory cortex. Hear Res. 76,
References 188–202.
Hendry, S.H.C., Schwark, H.D., Jones, E.G., Yan, J., 1987. Numbers and proportions of
Atencio, C.A., Schreiner, C.E., 2010a. Columnar connectivity and laminar processing GABA-immunoreactive neurons in different areas of monkey cerebral cortex. J.
in cat primary auditory cortex. PLoS One 5, e9521. Neurosci. 7, 1503–1519.
Atencio, C.A., Schreiner, C.E., 2010b. Laminar diversity of dynamic sound processing Holmgren, C., Harkany, T., Svennenfors, B., Zilberter, Y., 2003. Pyramidal cell com-
in cat primary auditory cortex. J. Neurophysiol. 103, 192–205. munication within local networks in layer 2/3 of rat neocortex. J. Physiol. 551,
Atzori, M., Lei, S., Evans, D.I., Kanold, P.O., Phillips-Tansey, E., McIntyre, O., McBain, 139–153.
C.J., 2001. Differential synaptic processing separates stationary from transient Huang, C.L., Winer, J.A., 2000. Auditory thalamocortical projections in the cat: lam-
inputs to the auditory cortex. Nat. Neurosci. 4, 1230–1237. inar and areal patterns of input. J. Comp. Neurol. 427, 302–331.
Barbour, D.L., Callaway, E.M., 2008. Excitatory local connections of superficial neu- Huetz, C., Philibert, B., Edeline, J.M., 2009. A spike-timing code for discriminat-
rons in rat auditory cortex. J. Neurosci. 28, 11174–11185. ing conspecific vocalizations in the thalamocortical system of anesthetized and
Bacci, A., Huguenard, J.R., 2006. Enhancement of spike-timing precision by autaptic awake guinea pigs. J. Neurosci. 29, 334–350.
transmission in neocortical inhibitory interneurons. Neuron 49, 119–130. Huggenberger, S., Vater, M., Deisz, R.A., 2009. Interlaminar differences of intrinsic
Benson, D.A., Teas, D.C., 1976. Single unit study of binaural interaction in the auditory properties of pyramidal neurons in the auditory cortex of mice. Cereb. Cortex.
cortex of the chinchilla. Brain Res. 103, 313–338. 19, 1008–1018.
Brandner, S., Redies, H., 1990. The projection from medial geniculate to field AI in Imig, T.J., Adrián, H.O., 1977. Binaural columns in the primary field (A1) of cat audi-
cat: organization in the isofrequency dimension. J. Neurosci. 10, 50–61. tory cortex. Brain Res. 138, 241–257.
Broicher, T., Bidmon, H.J., Kamuf, B., Coulon, P., Gorji, A., Pape, H.C., Speckmann, E.J., Imig, T.J., Brugge, J.F., 1978. Sources and terminations of callosal axons related to
Budde, T., 2010. Thalamic afferent activation of supragranular layers in auditory binaural and frequency maps in primary auditory cortex of the cat. J. Comp.
cortex in vitro: a voltage sensitive dye study. Neuroscience 165, 371–385. Neurol. 182, 637–660.
Brosch, M., Schreiner, C.E., 1997. Time course of forward masking tuning curves in Imig, T.J., Reale, R.A., 1981. Ipsilateral corticocortical projections related to binaural
cat primary auditory cortex. J. Neurophysiol. 77, 923–943. columns in cat primary auditory cortex. J. Comp. Neurol. 203, 1–14.
Brugge, J.F., Dubrovsky, N.A., Aitkin, L.M., Anderson, D.J., 1969. Sensitivity of single Kang, Y., Kaneko, T., Ohishi, H., Endo, K., Araki, T., 1994. Spatiotemporally differen-
neurons in auditory cortex of cat to binaural tonal stimulation; effects of varying tial inhibition of pyramidal cells in the cat motor cortex. J. Neurophysiol. 71,
interaural time and intensity. J. Neurophysiol. 32, 1005–1024. 280–293.
Brugge, J.F., Merzenich, M.M., 1973. Responses of neurons in auditory cortex of the Kaur, S., Lazar, R., Metherate, R., 2004. Intracortical pathways determine breadth of
macaque monkey to monaural and binaural stimulation. J. Neurophysiol. 36, subthreshold frequency receptive fields in primary auditory cortex. J. Neuro-
1138–1158. physiol. 91, 2551–2567.
2092 H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093

Kawaguchi, Y., Kondo, S., 2002. Parvalbumin, somatostatin and cholecystokinin as Richardson, R.J., Blundon, J.A., Bayazitov, I.T., Zakharenko, S.S., 2009. Connectivity
chemical markers for specific GABAergic interneuron types in the rat frontal patterns revealed by mapping of active inputs on dendrites of thalamorecipient
cortex. J. Neurocytol. 31, 277–287. neurons in the auditory cortex. J. Neurosci. 29, 6406–6417.
Kitzes, L.M., Wrege, K.S., Cassady, J.M., 1980. Patterns of responses of cortical cells Rose, H.J., Metherate, R., 2001. Thalamic stimulation largely elicits orthodromic,
to binaural stimulation. J. Comp. Neurol. 192, 455–472. rather than antidromic, cortical activation in an auditory thalamocortical slice.
Kitzes, L.M., Doherty, D., 1994. Influence of callosal activity on units in the auditory Neuroscience 106, 331–340.
cortex of ferret (Mustela putorius). J. Neurophysiol. 71, 1740–1751. Rose, H.J., Metherate, R., 2005. Auditory thalamocortical transmission is reliable and
Liu, B.H., Wu, G.K., Arbuckle, R., Tao, H.W., Zhang, L.I., 2007. Defining cortical fre- temporally precise. J. Neurophysiol. 94, 2019–2030.
quency tuning with recurrent excitatory circuitry. Nat. Neurosci. 10, 1594–1600. Rüttgers, K., Aschoff, A., Friauf, E., 1990. Commissural connections between the
Machens, C.K., Wehr, M.S., Zador, A.M., 2004. Linearity of cortical receptive fields auditory cortices of the rat. Brain Res. 509, 71–79.
measured with natural sounds. J. Neurosci. 24, 1089–1100. Sadagopan, S., Wang, X., 2010. Contribution of inhibition to stimulus selectivity in
Markram, H., Toledo-Rodriguez, M., Wang, Y., Gupta, A., Silberberg, G., Wu, C., primary auditory cortex of awake primates. J. Neurosci. 30, 7314–7325.
2004. Interneurons of the neocortical inhibitory system. Nat. Rev. Neurosci. 5, Sakata, S., Harris, K.D., 2009. Laminar structure of spontaneous and sensory-evoked
793–807. population activity in auditory cortex. Neuron 64, 404–418.
McMullen, N.T., Glaser, E.M., 1982. Morphology and laminar distribution of non- Sanchez-Vives, M.V., McCormick, D.A., 2000. Cellular and network mechanisms of
pyramidal neurons in the auditory cortex of the rabbit. J. Comp. Neurol. 208, rhythmic recurrent activity in neocortex. Nat. Neurosci. 3, 1027–1034.
85–106. Schreiner, C.E., 1992. Functional organization of the auditory cortex: maps and
McMullen, N.T., de Venecia, R.K., 1993. Thalamocortical patches in auditory neocor- mechanisms. Curr. Opin. Neurobiol. 2, 516–521.
tex. Brain Res. 620, 317–322. Semple, M.N., Kitzes, L.M., 1985. Single-unit responses in the inferior colliculus:
Metherate, R., Cruikshank, S.J., 1999. Thalamocortical inputs trigger a propagating different consequences of contralateral and ipsilateral auditory stimulation. J.
envelope of gamma-band activity in auditory cortex in vitro. Exp. Brain Res. 126, Neurophysiol. 53, 1467–1482.
160–174. Semple, M.N., Kitzes, L.M., 1993a. Binaural processing of sound pressure level in cat
Meyer, G., González-Hernández, T.H., Ferres-Torres, R., 1989. The spiny stellate neu- primary auditory cortex: evidence for a representation based on absolute levels
rons in layer IV of the human auditory cortex. A Golgi study. Neuroscience 33, rather than interaural level differences. J. Neurophysiol. 69, 449–461.
489–498. Semple, M.N., Kitzes, L.M., 1993b. Focal selectivity for binaural sound pressure level
Middlebrooks, J.C., Dykes, R.W., Merzenich, M.M., 1980. Binaural response-specific in cat primary auditory cortex: two-way intensity network tuning. J. Neurophys-
bands in primary auditory cortex (AI) of the cat: topographical organization iol. 69, 462–473.
orthogonal to isofrequency contours. Brain Res. 181, 31–48. Shapley, R., Hawken, M., Ringach, D.L., 2003. Dynamics of orientation selectivity in
Miles, R., 2000. Diversity in inhibition. Science 287 (5451), 244–246. the primary visual cortex and the importance of cortical inhibition. Neuron 38,
Miller, L.M., Escabí, M.A., Read, H.L., Schreiner, C.E., 2001. Functional convergence 689–699.
of response properties in the auditory thalamocortical system. Neuron 32, Shatz, C.J., 1977. Anatomy of interhemispheric connections in the visual system of
151–160. Boston Siamese and ordinary cats. J. Comp. Neurol. 173, 497–518.
Mitani, A., Shimokouchi, M., 1985. Neuronal connections in the primary audi- Shu, Y., Hasenstaub, A., Badoual, M., Bal, T., McCormick, D.A., 2003. Barrages of synap-
tory cortex: an electrophysiological study in the cat. J. Comp. Neurol. 235, tic activity control the gain and sensitivity of cortical neurons. J. Neurosci. 23,
417–429. 10388–10401.
Mitani, A., Shimokouchi, M., Itoh, K., Nomura, S., Kudo, M., Mizuno, N., 1985. Mor- Smith, P.H., Populin, L.C., 2001. Fundamental differences between the thalamocorti-
phology and laminar organization of electrophysiologically identified neurons cal recipient layers of the cat auditory and visual cortices. J. Comp. Neurol. 436,
in the primary auditory cortex in the cat. J. Comp. Neurol. 235, 430–447. 508–519.
Müller-Preuss, P., Ploog, D., 1981. Inhibition of auditory cortical neurons during Staiger, J.F., Schubert, D., Zuschratter, W., Kötter, R., Luhmann, H.J., Zilles, K., 2002.
phonation. Brain Res. 215, 61–76. Innervation of interneurons immunoreactive for VIP by intrinsically bursting
Ojima, H., Honda, C.N., Jones, E.G., 1991. Patterns of axon collateralization of identi- pyramidal cells and fast-spiking interneurons in infragranular layers of juvenile
fied supragranular pyramidal neurons in the cat auditory cortex. Cereb Cortex. rat neocortex. Eur. J. Neurosci. 16, 11–20.
1, 80–94. Staiger, J.F., Zuschratter, W., Luhmann, H.J., Schubert, D., 2009. Local circuits target-
Ojima, H., Honda, C.N., Jones, E.G., 1992. Characteristics of intracellularly injected ing parvalbumin-containing interneurons in layer IV of rat barrel cortex. Brain
infragranular pyramidal neurons in cat primary auditory cortex. Cereb. Cortex. Struct. Funct. 214, 1–13.
2, 197–216. Suga, N., 1977. Amplitude spectrum representation in the Doppler-shifted-CF pro-
Ojima, H., Murakami, K., 2002. Intracellular characterization of suppressive cessing area of the auditory cortex of the mustache bat. Science 196, 64–67.
responses in supragranular pyramidal neurons of cat primary auditory cortex Sutter, M.L., Schreiner, C.E., 1995. Topography of intensity tu ning in cat primary
in vivo. Cereb. Cortex. 12, 1079–1091. auditory cortex: single-neuron versus multiple-neuron recordings. J. Neuro-
Ojima, H., Taoka, M., Iriki, A., 2010. Adaptive changes in firing of primary auditory physiol. 73, 190–204.
cortical neurons following illumination shift from light to dark in freely moving Sutter, M.L., Schreiner, C.E., McLean, M., O’connor, K.N., Loftus, W.C., 1999. Organi-
guinea pigs. Cereb. Cortex. 20, 339–351. zation of inhibitory frequency receptive fields in cat primary auditory cortex. J.
Oswald, A.M., Schiff, M.L., Reyes, A.D., 2006. Synaptic mechanisms underlying audi- Neurophysiol. 82, 2358–2371.
tory processing. Curr. Opin. Neurobiol. 16, 371–376. Tan, A.Y., Atencio, C.A., Polley, D.B., Merzenich, M.M., Schreiner, C.E., 2007. Unbal-
Oswald, A.M., Reyes, A.D., 2008. Maturation of intrinsic and synaptic properties anced synaptic inhibition can create intensity-tuned auditory cortex neurons.
of layer 2/3 pyramidal neurons in mouse auditory cortex. J. Neurophysiol. 99, Neuroscience. 146, 449–462.
2998–3008. Tan, A.Y., Wehr, M., 2009. Balanced tone-evoked synaptic excitation and inhibition
Oswald, A.M., Doiron, B., Rinzel, J., Reyes, A.D., 2009. Spatial profile and differen- in mouse auditory cortex. Neuroscience 163, 1302–1315.
tial recruitment of GABAB modulate oscillatory activity in auditory cortex. J. Tan, A.Y., Zhang, L.I., Merzenich, M.M., Schreiner, C.E., 2004. Tone-evoked excita-
Neurosci. 29, 10321–10334. tory and inhibitory synaptic conductances of primary auditory cortex neurons.
Pfingst, B.E., O’Connor, T.A., 1981. Characteristics of neurons in auditory cortex of J. Neurophysiol. 92, 630–643.
monkeys performing a simple auditory task. J. Neurophysiol. 45, 16–34. Toyama, K., Matsunami, K., Kohno, T., Tokashiki, S., 1974. An intracellular study of
Phillips, D.P., Irvine, D.R., 1981. Responses of single neurons in physiologically neuronal organization in the visual cortex. Exp. Brain Res. 21, 45–66.
defined area AI of cat cerebral cortex: sensitivity to interaural intensity differ- Thomson, A.M., Lamy, C., 2007. Functional maps of neocortical local circuitry. Front.
ences. Hear Res. 4, 299–307. Neurosci. 1, 19–42.
Phillips, D.P., Irvine, D.R., 1983. Some features of binaural input to single neurons Valentine, P.A., Eggermont, J.J., 2001. Spontaneous burst-firing in three auditory
in physiologically defined area AI of cat cerebral cortex. J. Neurophysiol. 149, cortical fields: its relation to local field potentials and its effect on inter-area
383–395. cross-correlations. Hear Res. 154, 146–157.
Phillips, D.P., Cynader, M.S., 1985. Some neural mechanisms in the cat’s auditory cor- Valentine, P.A., Eggermont, J.J., 2003. Intracortical microstimulation induced
tex underlying sensitivity to combined tone and wide-spectrum noise stimuli. changes in spectral and temporal response properties in cat auditory cortex.
Hear Res. 18, 87–102. Hear Res. 183, 109–125.
Phillips, D.P., Orman, S.S., Musicant, A.D., Wilson, G.F., 1985. Neurons in the cat’s Volkov, I.O., Galazjuk, A.V., 1991. Formation of spike response to sound tones in
primary auditory cortex distinguished by their responses to tones and wide- cat auditory cortex neurons: interaction of excitatory and inhibitory effects.
spectrum noise. Hear Res. 18, 73–86. Neuroscience 43, 307–321.
Polley, D.B., Heiser, M.A., Blake, D.T., Schreiner, C.E., Merzenich, M.M., 2004. Associa- Volkov, I.O., Galazyuk, A.V., 1992. Peculiarities of inhibition in cat auditory cor-
tive learning shapes the neural code for stimulus magnitude in primary auditory tex neurons evoked by tonal stimuli of various durations. Exp. Brain Res. 91,
cortex. Proc. Natl. Acad. Sci. U. S. A. 101, 16351–16356. 115–120.
Polley, D.B., Steinberg, E.E., Merzenich, M.M., 2006. Perceptual learning directs audi- Wallace, M.N., Harper, M.S., 1997. Callosal connections of the ferret primary auditory
tory cortical map reorganization through top-down influences. J Neurosci. 26, cortex. Exp. Brain Res. 116, 367–374.
4970–4982. Wang, J., Caspary, D., Salvi, R.J., 2000. GABA-A antagonist causes dramatic expansion
Prieto, J.J., Peterson, B.A., Winer, J.A., 1994. Morphology and spatial distribution of of tuning in primary auditory cortex. Neuroreport 11, 1137–1140.
GABAergic neurons in cat primary auditory cortex (AI). J. Comp. Neurol. 344, Wang, J., McFadden, S.L., Caspary, D., Salvi, R., 2002. Gamma-aminobutyric acid
349–382. circuits shape response properties of auditory cortex neurons. Brain Res. 944,
Priebe, N.J., Ferster, D., 2008. Inhibition, spike threshold, and stimulus selectivity in 219–231.
primary visual cortex. Neuron 57, 482–497. Wehr, M., Zador, A.M., 2003. Balanced inhibition underlies tuning and sharpens spike
timing in auditory cortex. Nature 426, 442–446.
H. Ojima / Neuroscience and Biobehavioral Reviews 35 (2011) 2084–2093 2093

Wehr, M., Zador, A.M., 2005. Synaptic mechanisms of forward suppression in rat Ye, C.-Q., Poo, M.-M., Dan, Y., Zhang, X.-N., 2010. Synaptic mechanisms of direction
auditory cortex. Neuron 47, 437–445. selectivity in primary auditory cortex. J. Neurosci. 30, 1861–1868.
Winer, J.A., 1984. Anatomy of layer IV in cat primary auditory cortex (AI). J. Comp. Zhou, Y., Liu, B.H., Wu, G.K., Kim, Y.J., Xiao, Z., Tao, H.W., Zhang, L.I., 2010. Pre-
Neurol. 224, 535–567. ceding inhibition silences layer 6 neurons in auditory cortex. Neuron 65,
Wu, G.K., Arbuckle, R., Liu, B.H., Tao, H.W., Zhang, L.I., 2008. Lateral sharpening 706–717.
of cortical frequency tuning by approximately balanced inhibition. Neuron 58, Zhang, L.I., Tan, A.Y., Schreiner, C.E., Merzenich, M.M., 2003. Topography and synaptic
132–143. shaping of direction selectivity in primary auditory cortex. Nature 424, 201–205.
Wu, G.K., Li, P., Tao, H.W., Zhang, L.I., 2006. Nonmonotonic synaptic excitation and Zhang, J., Nakamoto, K.T., Kitzes, L.M., 2004. Binaural interaction revisited in the cat
imbalanced inhibition underlying cortical intensity tuning. Neuron 52, 705–715. primary auditory cortex. J. Neurophysiol. 91, 101–117.

You might also like